Courcelle's theorem

Last updated

In the study of graph algorithms, Courcelle's theorem is the statement that every graph property definable in the monadic second-order logic of graphs can be decided in linear time on graphs of bounded treewidth. [1] [2] [3] The result was first proved by Bruno Courcelle in 1990 [4] and independently rediscovered by Borie, Parker & Tovey (1992). [5] It is considered the archetype of algorithmic meta-theorems. [6] [7]

Contents

Formulations

Vertex sets

In one variation of monadic second-order graph logic known as MSO1, the graph is described by a set of vertices and a binary adjacency relation , and the restriction to monadic logic means that the graph property in question may be defined in terms of sets of vertices of the given graph, but not in terms of sets of edges, or sets of tuples of vertices.

As an example, the property of a graph being colorable with three colors (represented by three sets of vertices , , and ) may be defined by the monadic second-order formula

with the naming convention that uppercase variables denote sets of vertices and lowercase variables denote individual vertices (so that an explicit declaration of which is which can be omitted from the formula). The first part of this formula ensures that the three color classes cover all the vertices of the graph, and the rest ensures that they each form an independent set. (It would also be possible to add clauses to the formula to ensure that the three color classes are disjoint, but this makes no difference to the result.) Thus, by Courcelle's theorem, 3-colorability of graphs of bounded treewidth may be tested in linear time.

For this variation of graph logic, Courcelle's theorem can be extended from treewidth to clique-width: for every fixed MSO1 property , and every fixed bound on the clique-width of a graph, there is a linear-time algorithm for testing whether a graph of clique-width at most has property . [8] The original formulation of this result required the input graph to be given together with a construction proving that it has bounded clique-width, but later approximation algorithms for clique-width removed this requirement. [9]

Edge sets

Courcelle's theorem may also be used with a stronger variation of monadic second-order logic known as MSO2. In this formulation, a graph is represented by a set V of vertices, a set E of edges, and an incidence relation between vertices and edges. This variation allows quantification over sets of vertices or edges, but not over more complex relations on tuples of vertices or edges.

For instance, the property of having a Hamiltonian cycle may be expressed in MSO2 by describing the cycle as a set of edges that includes exactly two edges incident to each vertex, such that every nonempty proper subset of vertices has an edge in the putative cycle with exactly one endpoint in the subset. However, Hamiltonicity cannot be expressed in MSO1. [10]

Labeled graphs

It is possible to apply the same results to graphs in which the vertices or edges have labels from a fixed finite set, either by augmenting the graph logic to incorporate predicates describing the labels, or by representing the labels by unquantified vertex set or edge set variables. [11]

Modular congruences

Another direction for extending Courcelle's theorem concerns logical formulas that include predicates for counting the size of the test. In this context, it is not possible to perform arbitrary arithmetic operations on set sizes, nor even to test whether two sets have the same size. However, MSO1 and MSO2 can be extended to logics called CMSO1 and CMSO2, that include for every two constants q and r a predicate which tests whether the cardinality of set S is congruent to r modulo q. Courcelle's theorem can be extended to these logics. [4]

Decision versus optimization

As stated above, Courcelle's theorem applies primarily to decision problems: does a graph have a property or not. However, the same methods also allow the solution to optimization problems in which the vertices or edges of a graph have integer weights, and one seeks the minimum or maximum weight vertex set that satisfies a given property, expressed in second-order logic. These optimization problems can be solved in linear time on graphs of bounded clique-width. [8] [11]

Space complexity

Rather than bounding the time complexity of an algorithm that recognizes an MSO property on bounded-treewidth graphs, it is also possible to analyze the space complexity of such an algorithm; that is, the amount of memory needed above and beyond the size of the input itself (which is assumed to be represented in a read-only way so that its space requirements cannot be put to other purposes). In particular, it is possible to recognize the graphs of bounded treewidth, and any MSO property on these graphs, by a deterministic Turing machine that uses only logarithmic space. [12]

Proof strategy and complexity

The typical approach to proving Courcelle's theorem involves the construction of a finite bottom-up tree automaton that acts on the tree decompositions of the given graph. [6]

In more detail, two graphs G1 and G2, each with a specified subset T of vertices called terminals, may be defined to be equivalent with respect to an MSO formula F if, for all other graphs H whose intersection with G1 and G2 consists only of vertices in T, the two graphs G1  H and G2  H behave the same with respect to F: either they both model F or they both do not model F. This is an equivalence relation, and it can be shown by induction on the length of F that (when the sizes of T and F are both bounded) it has finitely many equivalence classes. [13]

A tree decomposition of a given graph G consists of a tree and, for each tree node, a subset of the vertices of G called a bag. It must satisfy two properties: for each vertex v of G, the bags containing v must be associated with a contiguous subtree of the tree, and for each edge uv of G, there must be a bag containing both u and v. The vertices in a bag can be thought of as the terminals of a subgraph of G, represented by the subtree of the tree decomposition descending from that bag. When G has bounded treewidth, it has a tree decomposition in which all bags have bounded size, and such a decomposition can be found in fixed-parameter tractable time. [14] Moreover, it is possible to choose this tree decomposition so that it forms a binary tree, with only two child subtrees per bag. Therefore, it is possible to perform a bottom-up computation on this tree decomposition, computing an identifier for the equivalence class of the subtree rooted at each bag by combining the edges represented within the bag with the two identifiers for the equivalence classes of its two children. [15]

The size of the automaton constructed in this way is not an elementary function of the size of the input MSO formula. This non-elementary complexity is necessary, in the sense that (unless P = NP) it is not possible to test MSO properties on trees in a time that is fixed-parameter tractable with an elementary dependence on the parameter. [16]

Bojańczyk-Pilipczuk's theorem

The proofs of Courcelle's theorem show a stronger result: not only can every (counting) monadic second-order property be recognized in linear time for graphs of bounded treewidth, but also it can be recognized by a finite-state tree automaton. Courcelle conjectured a converse to this: if a property of graphs of bounded treewidth is recognized by a tree automaton, then it can be defined in counting monadic second-order logic. In 1998 Lapoire (1998), claimed a resolution of the conjecture. [17] However, the proof is widely regarded as unsatisfactory. [18] [19] Until 2016, only a few special cases were resolved: in particular, the conjecture has been proved for graphs of treewidth at most three, [20] for k-connected graphs of treewidth k, for graphs of constant treewidth and chordality, and for k-outerplanar graphs. The general version of the conjecture was finally proved by Mikołaj Bojańczyk and Michał Pilipczuk. [21]

Moreover, for Halin graphs (a special case of treewidth three graphs) counting is not needed: for these graphs, every property that can be recognized by a tree automaton can also be defined in monadic second-order logic. The same is true more generally for certain classes of graphs in which a tree decomposition can itself be described in MSOL. However, it cannot be true for all graphs of bounded treewidth, because in general counting adds extra power over monadic second-order logic without counting. For instance, the graphs with an even number of vertices can be recognized using counting, but not without. [19]

Satisfiability and Seese's theorem

The satisfiability problem for a formula of monadic second-order logic is the problem of determining whether there exists at least one graph (possibly within a restricted family of graphs) for which the formula is true. For arbitrary graph families, and arbitrary formulas, this problem is undecidable. However, satisfiability of MSO2 formulas is decidable for the graphs of bounded treewidth, and satisfiability of MSO1 formulas is decidable for graphs of bounded clique-width. The proof involves building a tree automaton for the formula and then testing whether the automaton has an accepting path.

As a partial converse, Seese (1991) proved that, whenever a family of graphs has a decidable MSO2 satisfiability problem, the family must have bounded treewidth. The proof is based on a theorem of Robertson and Seymour that the families of graphs with unbounded treewidth have arbitrarily large grid minors. [22] Seese also conjectured that every family of graphs with a decidable MSO1 satisfiability problem must have bounded clique-width; this has not been proven, but a weakening of the conjecture that replaces MSO1 by CMSO1 is true. [23]

Applications

Grohe (2001) used Courcelle's theorem to show that computing the crossing number of a graph G is fixed-parameter tractable with a quadratic dependence on the size of G, improving a cubic-time algorithm based on the Robertson–Seymour theorem. An additional later improvement to linear time by Kawarabayashi & Reed (2007) follows the same approach. If the given graph G has small treewidth, Courcelle's theorem can be applied directly to this problem. On the other hand, if G has large treewidth, then it contains a large grid minor, within which the graph can be simplified while leaving the crossing number unchanged. Grohe's algorithm performs these simplifications until the remaining graph has a small treewidth, and then applies Courcelle's theorem to solve the reduced subproblem. [24] [25]

Gottlob & Lee (2007) observed that Courcelle's theorem applies to several problems of finding minimum multi-way cuts in a graph, when the structure formed by the graph and the set of cut pairs has bounded treewidth. As a result they obtain a fixed-parameter tractable algorithm for these problems, parameterized by a single parameter, treewidth, improving previous solutions that had combined multiple parameters. [26]

In computational topology, Burton & Downey (2014) extend Courcelle's theorem from MSO2 to a form of monadic second-order logic on simplicial complexes of bounded dimension that allows quantification over simplices of any fixed dimension. As a consequence, they show how to compute certain quantum invariants of 3-manifolds as well as how to solve certain problems in discrete Morse theory efficiently, when the manifold has a triangulation (avoiding degenerate simplices) whose dual graph has small treewidth. [27]

Methods based on Courcelle's theorem have also been applied to database theory, [28] knowledge representation and reasoning, [29] automata theory, [30] and model checking. [31]

Related Research Articles

<span class="mw-page-title-main">Tree decomposition</span> Mapping of a graph into a tree

In graph theory, a tree decomposition is a mapping of a graph into a tree that can be used to define the treewidth of the graph and speed up solving certain computational problems on the graph.

In graph theory, an undirected graph H is called a minor of the graph G if H can be formed from G by deleting edges, vertices and by contracting edges.

<span class="mw-page-title-main">Chordal graph</span> Graph where all long cycles have a chord

In the mathematical area of graph theory, a chordal graph is one in which all cycles of four or more vertices have a chord, which is an edge that is not part of the cycle but connects two vertices of the cycle. Equivalently, every induced cycle in the graph should have exactly three vertices. The chordal graphs may also be characterized as the graphs that have perfect elimination orderings, as the graphs in which each minimal separator is a clique, and as the intersection graphs of subtrees of a tree. They are sometimes also called rigid circuit graphs or triangulated graphs: a chordal completion of a graph is typically called a triangulation of that graph.

<span class="mw-page-title-main">Cograph</span> Graph formed by complementation and disjoint union

In graph theory, a cograph, or complement-reducible graph, or P4-free graph, is a graph that can be generated from the single-vertex graph K1 by complementation and disjoint union. That is, the family of cographs is the smallest class of graphs that includes K1 and is closed under complementation and disjoint union.

<span class="mw-page-title-main">Monochromatic triangle</span>

In graph theory and theoretical computer science, the monochromatic triangle problem is an algorithmic problem on graphs, in which the goal is to partition the edges of a given graph into two triangle-free subgraphs. It is NP-complete but fixed-parameter tractable on graphs of bounded treewidth.

In graph theory, the treewidth of an undirected graph is an integer number which specifies, informally, how far the graph is from being a tree. The smallest treewidth is 1; the graphs with treewidth 1 are exactly the trees and the forests. The graphs with treewidth at most 2 are the series–parallel graphs. The maximal graphs with treewidth exactly k are called k-trees, and the graphs with treewidth at most k are called partial k-trees. Many other well-studied graph families also have bounded treewidth.

In graph theory, a path decomposition of a graph G is, informally, a representation of G as a "thickened" path graph, and the pathwidth of G is a number that measures how much the path was thickened to form G. More formally, a path-decomposition is a sequence of subsets of vertices of G such that the endpoints of each edge appear in one of the subsets and such that each vertex appears in a contiguous subsequence of the subsets, and the pathwidth is one less than the size of the largest set in such a decomposition. Pathwidth is also known as interval thickness, vertex separation number, or node searching number.

<span class="mw-page-title-main">Strong product of graphs</span> Binary operation in graph theory

In graph theory, the strong product is a way of combining two graphs to make a larger graph. Two vertices are adjacent in the strong product when they come from pairs of vertices in the factor graphs that are either adjacent or identical. The strong product is one of several different graph product operations that have been studied in graph theory. The strong product of any two graphs can be constructed as the union of two other products of the same two graphs, the Cartesian product of graphs and the tensor product of graphs.

<span class="mw-page-title-main">Hosoya index</span> Number of matchings in a graph

The Hosoya index, also known as the Z index, of a graph is the total number of matchings in it. The Hosoya index is always at least one, because the empty set of edges is counted as a matching for this purpose. Equivalently, the Hosoya index is the number of non-empty matchings plus one. The index is named after Haruo Hosoya. It is used as a topological index in chemical graph theory.

In mathematical logic, monadic second-order logic (MSO) is the fragment of second-order logic where the second-order quantification is limited to quantification over sets. It is particularly important in the logic of graphs, because of Courcelle's theorem, which provides algorithms for evaluating monadic second-order formulas over graphs of bounded treewidth. It is also of fundamental importance in automata theory, where the Büchi–Elgot–Trakhtenbrot theorem gives a logical characterization of the regular languages.

<span class="mw-page-title-main">Clique-width</span> Measure of graph complexity

In graph theory, the clique-width of a graph G is a parameter that describes the structural complexity of the graph; it is closely related to treewidth, but unlike treewidth it can be small for dense graphs. It is defined as the minimum number of labels needed to construct G by means of the following 4 operations :

  1. Creation of a new vertex v with label i (denoted by i(v))
  2. Disjoint union of two labeled graphs G and H (denoted by )
  3. Joining by an edge every vertex labeled i to every vertex labeled j (denoted by η(i,j)), where ij
  4. Renaming label i to label j (denoted by ρ(i,j))
<span class="mw-page-title-main">Branch-decomposition</span> Hierarchical clustering of graph edges

In graph theory, a branch-decomposition of an undirected graph G is a hierarchical clustering of the edges of G, represented by an unrooted binary tree T with the edges of G as its leaves. Removing any edge from T partitions the edges of G into two subgraphs, and the width of the decomposition is the maximum number of shared vertices of any pair of subgraphs formed in this way. The branchwidth of G is the minimum width of any branch-decomposition of G.

<span class="mw-page-title-main">Distance-hereditary graph</span> Graph whose induced subgraphs preserve distance

In graph theory, a branch of discrete mathematics, a distance-hereditary graph is a graph in which the distances in any connected induced subgraph are the same as they are in the original graph. Thus, any induced subgraph inherits the distances of the larger graph.

<span class="mw-page-title-main">Clique-sum</span> Gluing graphs at complete subgraphs

In graph theory, a branch of mathematics, a clique sum is a way of combining two graphs by gluing them together at a clique, analogous to the connected sum operation in topology. If two graphs G and H each contain cliques of equal size, the clique-sum of G and H is formed from their disjoint union by identifying pairs of vertices in these two cliques to form a single shared clique, and then deleting all the clique edges or possibly deleting some of the clique edges. A k-clique-sum is a clique-sum in which both cliques have exactly k vertices. One may also form clique-sums and k-clique-sums of more than two graphs, by repeated application of the clique-sum operation.

In graph theory, the tree-depth of a connected undirected graph is a numerical invariant of , the minimum height of a Trémaux tree for a supergraph of . This invariant and its close relatives have gone under many different names in the literature, including vertex ranking number, ordered chromatic number, and minimum elimination tree height; it is also closely related to the cycle rank of directed graphs and the star height of regular languages. Intuitively, where the treewidth of a graph measures how far it is from being a tree, this parameter measures how far a graph is from being a star.

In graph theory, a Trémaux tree of an undirected graph is a type of spanning tree, generalizing depth-first search trees. They are defined by the property that every edge of connects an ancestor–descendant pair in the tree. Trémaux trees are named after Charles Pierre Trémaux, a 19th-century French author who used a form of depth-first search as a strategy for solving mazes. They have also been called normal spanning trees, especially in the context of infinite graphs.

<span class="mw-page-title-main">Bramble (graph theory)</span> Method of graph decomposition

In graph theory, a bramble for an undirected graph G is a family of connected subgraphs of G that all touch each other: for every pair of disjoint subgraphs, there must exist an edge in G that has one endpoint in each subgraph. The order of a bramble is the smallest size of a hitting set, a set of vertices of G that has a nonempty intersection with each of the subgraphs. Brambles may be used to characterize the treewidth of G.

In the mathematical fields of graph theory and finite model theory, the logic of graphs deals with formal specifications of graph properties using sentences of mathematical logic. There are several variations in the types of logical operation that can be used in these sentences. The first-order logic of graphs concerns sentences in which the variables and predicates concern individual vertices and edges of a graph, while monadic second-order graph logic allows quantification over sets of vertices or edges. Logics based on least fixed point operators allow more general predicates over tuples of vertices, but these predicates can only be constructed through fixed-point operators, restricting their power.

<i>k</i>-outerplanar graph

In graph theory, a k-outerplanar graph is a planar graph that has a planar embedding in which the vertices belong to at most concentric layers. The outerplanarity index of a planar graph is the minimum value of for which it is -outerplanar.

In mathematics, S2S is the monadic second order theory with two successors. It is one of the most expressive natural decidable theories known, with many decidable theories interpretable in S2S. Its decidability was proved by Rabin in 1969.

References

  1. Eger, Steffen (2008), Regular Languages, Tree Width, and Courcelle's Theorem: An Introduction, VDM Publishing, ISBN   9783639076332 .
  2. Courcelle, Bruno; Engelfriet, Joost (2012), Graph Structure and Monadic Second-Order Logic: A Language-Theoretic Approach (PDF), Encyclopedia of Mathematics and its Applications, vol. 138, Cambridge University Press, ISBN   9781139644006, Zbl   1257.68006 .
  3. Downey, Rodney G.; Fellows, Michael R. (2013), "Chapter 13: Courcelle's theorem", Fundamentals of parameterized complexity, Texts in Computer Science, London: Springer, pp. 265–278, CiteSeerX   10.1.1.456.2729 , doi:10.1007/978-1-4471-5559-1, ISBN   978-1-4471-5558-4, MR   3154461, S2CID   23517218 .
  4. 1 2 Courcelle, Bruno (1990), "The monadic second-order logic of graphs. I. Recognizable sets of finite graphs", Information and Computation, 85 (1): 12–75, doi: 10.1016/0890-5401(90)90043-H , MR   1042649, Zbl   0722.03008
  5. Borie, Richard B.; Parker, R. Gary; Tovey, Craig A. (1992), "Automatic generation of linear-time algorithms from predicate calculus descriptions of problems on recursively constructed graph families", Algorithmica, 7 (5–6): 555–581, doi:10.1007/BF01758777, MR   1154588, S2CID   22623740 .
  6. 1 2 Kneis, Joachim; Langer, Alexander (2009), "A practical approach to Courcelle's theorem", Electronic Notes in Theoretical Computer Science, 251: 65–81, doi: 10.1016/j.entcs.2009.08.028 .
  7. Lampis, Michael (2010), "Algorithmic meta-theorems for restrictions of treewidth", in de Berg, Mark; Meyer, Ulrich (eds.), Proc. 18th Annual European Symposium on Algorithms, Lecture Notes in Computer Science, vol. 6346, Springer, pp. 549–560, doi:10.1007/978-3-642-15775-2_47, Zbl   1287.68078 .
  8. 1 2 Courcelle, B.; Makowsky, J. A.; Rotics, U. (2000), "Linear time solvable optimization problems on graphs of bounded clique-width", Theory of Computing Systems, 33 (2): 125–150, CiteSeerX   10.1.1.414.1845 , doi:10.1007/s002249910009, MR   1739644, S2CID   15402031, Zbl   1009.68102 .
  9. Oum, Sang-il; Seymour, Paul (2006), "Approximating clique-width and branch-width", Journal of Combinatorial Theory , Series B, 96 (4): 514–528, doi: 10.1016/j.jctb.2005.10.006 , MR   2232389 .
  10. Courcelle & Engelfriet (2012), Proposition 5.13, p. 338.
  11. 1 2 Arnborg, Stefan; Lagergren, Jens; Seese, Detlef (1991), "Easy problems for tree-decomposable graphs", Journal of Algorithms, 12 (2): 308–340, CiteSeerX   10.1.1.12.2544 , doi:10.1016/0196-6774(91)90006-K, MR   1105479 .
  12. Elberfeld, Michael; Jakoby, Andreas; Tantau, Till (October 2010), "Logspace Versions of the Theorems of Bodlaender and Courcelle" (PDF), Proc. 51st Annual IEEE Symposium on Foundations of Computer Science (FOCS 2010) , pp. 143–152, doi:10.1109/FOCS.2010.21, S2CID   1820251 .
  13. Downey & Fellows (2013), Theorem 13.1.1, p. 266.
  14. Downey & Fellows (2013), Section 10.5: Bodlaender's theorem, pp. 195–203.
  15. Downey & Fellows (2013), Section 12.6: Tree automata, pp. 237–247.
  16. Frick, Markus; Grohe, Martin (2004), "The complexity of first-order and monadic second-order logic revisited", Annals of Pure and Applied Logic, 130 (1–3): 3–31, CiteSeerX   10.1.1.104.8429 , doi: 10.1016/j.apal.2004.01.007 , MR   2092847 .
  17. Lapoire, Denis (1998), "Recognizability equals monadic second-order definability for sets of graphs of bounded tree-width", STACS 98: 15th Annual Symposium on Theoretical Aspects of Computer Science Paris, France, February 27, 1998, Proceedings, vol. 1373, pp. 618–628, Bibcode:1998LNCS.1373..618L, CiteSeerX   10.1.1.22.7805 , doi:10.1007/bfb0028596 .
  18. Courcelle, B.; Engelfriet., J. (2012), "Graph Structure and Monadic Second Order Logic -- A Language-Theoretic Approach", Encyclopedia of mathematics and its applications, vol. 138, Cambridge University Press.
  19. 1 2 Jaffke, Lars; Bodlaender, Hans L. (2015), MSOL-definability equals recognizability for Halin graphs and bounded degree k-outerplanar graphs, arXiv: 1503.01604 , Bibcode:2015arXiv150301604J .
  20. Kaller, D. (2000), "Definability equals recognizability of partial 3-trees and k-connected partial k-trees", Algorithmica, 27 (3): 348–381, doi:10.1007/s004530010024, S2CID   39798483 .
  21. Bojańczyk, Mikołaj; Pilipczuk, Michał (2016), "Definability equals recognizability for graphs of bounded treewidth", Proceedings of the 31st Annual ACM/IEEE Symposium on Logic in Computer Science (LICS 2016), pp. 407–416, arXiv: 1605.03045 , doi:10.1145/2933575.2934508, S2CID   1213054 .
  22. Seese, D. (1991), "The structure of the models of decidable monadic theories of graphs", Annals of Pure and Applied Logic, 53 (2): 169–195, doi:10.1016/0168-0072(91)90054-P, MR   1114848 .
  23. Courcelle, Bruno; Oum, Sang-il (2007), "Vertex-minors, monadic second-order logic, and a conjecture by Seese" (PDF), Journal of Combinatorial Theory , Series B, 97 (1): 91–126, doi: 10.1016/j.jctb.2006.04.003 , MR   2278126 .
  24. Grohe, Martin (2001), "Computing crossing numbers in quadratic time", Proceedings of the Thirty-Third Annual ACM Symposium on Theory of Computing (STOC '01) , pp. 231–236, arXiv: cs/0009010 , doi:10.1145/380752.380805, S2CID   724544 .
  25. Kawarabayashi, Ken-ichi; Reed, Bruce (2007), "Computing crossing number in linear time", Proceedings of the Thirty-Ninth Annual ACM Symposium on Theory of Computing (STOC '07) , pp. 382–390, doi:10.1145/1250790.1250848, S2CID   13000831 .
  26. Gottlob, Georg; Lee, Stephanie Tien (2007), "A logical approach to multicut problems", Information Processing Letters , 103 (4): 136–141, doi:10.1016/j.ipl.2007.03.005, MR   2330167 .
  27. Burton, Benjamin A.; Downey, Rodney G. (2014), Courcelle's theorem for triangulations, arXiv: 1403.2926 , Bibcode:2014arXiv1403.2926B . Short communication, International Congress of Mathematicians, 2014.
  28. Grohe, Martin; Mariño, Julian (1999), "Definability and descriptive complexity on databases of bounded tree-width", Database Theory — ICDT'99: 7th International Conference Jerusalem, Israel, January 10–12, 1999, Proceedings, Lecture Notes in Computer Science, vol. 1540, pp. 70–82, CiteSeerX   10.1.1.52.2984 , doi:10.1007/3-540-49257-7_6 .
  29. Gottlob, Georg; Pichler, Reinhard; Wei, Fang (January 2010), "Bounded treewidth as a key to tractability of knowledge representation and reasoning", Artificial Intelligence, 174 (1): 105–132, doi: 10.1016/j.artint.2009.10.003 .
  30. Madhusudan, P.; Parlato, Gennaro (2011), "The Tree Width of Auxiliary Storage", Proceedings of the 38th Annual ACM SIGPLAN-SIGACT Symposium on Principles of Programming Languages (POPL '11) (PDF), SIGPLAN Notices, vol. 46, pp. 283–294, doi:10.1145/1926385.1926419, S2CID   6976816
  31. Obdržálek, Jan (2003), "Fast mu-calculus model checking when tree-width is bounded", Computer Aided Verification: 15th International Conference, CAV 2003, Boulder, CO, USA, July 8-12, 2003, Proceedings, Lecture Notes in Computer Science, vol. 2725, pp. 80–92, CiteSeerX   10.1.1.2.4843 , doi:10.1007/978-3-540-45069-6_7 .